Water incorporation in omphacite: concentrations and...

28
Skogby et al. 1 Cover page 1 Water incorporation in omphacite: concentrations and compositional 2 relations in ultrahigh-pressure eclogites from Pohorje, Eastern Alps 3 4 Running title: Water incorporation in omphacite 5 Plan: 6 Abstract 7 Introduction 8 Geological background and petrography 9 Experimental methods 10 Infrared spectroscopy 11 Mössbauer spectroscopy 12 Electron microprobe analysis 13 Results 14 Discussion 15 Acknowledgements 16 References 17 18 Corresponding author: 19 Henrik Skogby 20 Swedish Museum of Natural History 21 Department of Geosciences 22 Box 50007 23 SE-104 05 Stockholm 24 Sweden 25 e-mail: [email protected]. 26 phone: +46 8 51954043 27 28 29 30

Transcript of Water incorporation in omphacite: concentrations and...

  • Skogby et al.

    1

    Cover page 1

    Water incorporation in omphacite: concentrations and compositional 2

    relations in ultrahigh-pressure eclogites from Pohorje, Eastern Alps 3

    4

    Running title: Water incorporation in omphacite 5

    Plan: 6

    Abstract 7

    Introduction 8

    Geological background and petrography 9

    Experimental methods 10

    Infrared spectroscopy 11

    Mössbauer spectroscopy 12

    Electron microprobe analysis 13

    Results 14

    Discussion 15

    Acknowledgements 16

    References 17

    18

    Corresponding author: 19

    Henrik Skogby 20

    Swedish Museum of Natural History 21

    Department of Geosciences 22

    Box 50007 23

    SE-104 05 Stockholm 24

    Sweden 25

    e-mail: [email protected]. 26

    phone: +46 8 51954043 27

    28

    29

    30

    mailto:[email protected]

  • Skogby et al.

    2

    31

    Title page 32

    33

    Water incorporation in omphacite: concentrations and compositional 34

    relations in ultrahigh-pressure eclogites from Pohorje, Eastern Alps 35

    36

    Henrik Skogby1*

    , Marian Janák2, and Igor Broska

    2 37

    38 1Swedish Museum of Natural History, Department of Geosciences, Box 50007, SE-104 05 39

    Stockholm, Sweden. 40

    * Corresponding author, e-mail: [email protected]. 41 2Earth Science Institute, Slovak Academy of Sciences, Dúbravská 9, Box 106, 840 05 42

    Bratislava, Slovak Republic. 43

    44

    45

  • Skogby et al.

    3

    46

    Abstract: Omphacite in ultra-high pressure (UHP) eclogites from the Pohorje Mountains in 47

    Slovenia, south-eastern Alps, has been investigated by electron microprobe (EMP), Infra-Red 48

    (IR) and Mössbauer spectroscopy to determine OH concentrations and related incorporation 49

    mechanisms. Results from polarized IR measurements reveal high contents of structurally-50

    bound OH, varying from 530 to 870 ppm H2O. Characterization of omphacite composition by 51

    EMP analysis and Mössbauer spectroscopy shows that all samples contain vacancies at the 52

    M2 position, which can be expressed as a Ca-Eskola component (Ca0.5□0.5AlSi2O6). The 53

    amount of the Ca-Eskola component displays a positive correlation with the OH 54

    concentration, which confirms results from previous studies. The occurrence of precipitated 55

    quartz rods in some samples indicates that primary omphacite contained a larger Ca-Eskola 56

    component. Extrapolation of the observed trend-line for the Ca-Eskola and OH contents point 57

    to an original OH concentration around 1500 ppm H2O for these samples. The high water 58

    contents observed in omphacite are considered to be linked to the UHP origin of the eclogite 59

    rocks. 60

    61

    Key words: Omphacite; Ca-Eskola; eclogite; UHP; infrared spectroscopy; Mössbauer 62

    spectroscopy. 63

    64

    65

    Introduction 66

    67

    The pyroxenes are one of the nominally anhydrous mineral groups that have been shown to 68

    generally contain low but significant contents of hydrogen, bound in their crystal structures as 69

    OH- ions. The amounts of hydrogen incorporated in pyroxene are considered to be related to 70

    both intrinsic factors, such as sample composition and defect chemistry, as well as to external 71

  • Skogby et al.

    4

    ones such as T, Ptot and PH2O conditions, thereby reflecting the equilibration environment. 72

    During recent years, hydrogen contents in pyroxenes have been extensively explored to gain 73

    information from a range of geological environments, including water storage mechanisms 74

    and circulation in Earth’s mantle (e.g. Ingrin & Skogby, 2000; Bolfan-Casanova, 2005; 75

    Peslier, 2010) as well as magmatic water contents in volcanic systems (e.g. Wade et al., 2008; 76

    O’Leary et al., 2010; Weis et al., 2015). 77

    The modes of hydrogen accommodation in clinopyroxene are not yet fully understood. 78

    However, most incorporation models (e.g. Andrut et al., 2007) have suggested that OH 79

    preferentially occurs on the underbonded O2 position, with charge compensation provided 80

    either by vacancies at the M2 position or charge-deficient substitutions such as Al replacing 81

    Si at the tetrahedral site. Among the pyroxenes, omphacite has been shown to frequently 82

    accommodate relatively high concentrations of OH (Skogby et al., 1990, Smyth et al., 1991; 83

    Katayama & Nakashima, 2003; Katayama et al., 2006; Koch-Müller et al., 2004; 2007; 84

    Konzett et al., 2008a), approaching close to 1000 ppm H2O. These high contents appear to be 85

    coupled to formation of structural vacancies via the so-called Ca-Eskola (CaEs) component, 86

    Ca0.5□0.5AlSi2O6, which has been shown to be essential for OH incorporation in omphacite. 87

    Such a relation was first reported by Smyth et al. (1991) for a series of jadeite-rich 88

    omphacites from South-African kimberlites, where the amount of M2 vacancies was shown to 89

    be correlated with the absorbance of the OH-bands in FTIR spectra, and hence water 90

    concentration. This trend was further supported by Katayama & Nakashima (2003), who also 91

    found a positive correlation for OH absorbance and calculated CaEs components in a series of 92

    clinopyroxenes from deeply subducted crust occurrences in the Kokchetav UHP metamorphic 93

    terrane. In an experimental study of hydroxyl solubility in synthetic jadeite and Na-rich 94

    clinopyroxene, Bromiley & Keppler (2004) concluded that solid solutions of jadeite with 95

    diopside and in particular CaEs components lead to a drastic increase of water solubility, and 96

  • Skogby et al.

    5

    that these compositional effects were substantially larger than those imposed by pressure and 97

    temperature. Additional support for the importance of the CaEs component was provided by 98

    Koch-Müller et al. (2004) for a series of omphacites from Yakutian kimberlite pipes. 99

    Even though the studies mentioned above have clearly demonstrated that the OH 100

    content in omphacite is frequently correlated with the CaEs component, a full characterization 101

    of the amount of CaEs has generally not been established. This has been caused by lack of Fe 102

    valence determination, which is needed to calculate an appropriate CaEs (and vacancy) 103

    component, and has lead to underestimation and often negative values of the CaEs component 104

    when all Fe has been assumed as Fe2+

    . In order to increase the understanding of OH 105

    incorporation in omphacite and the role of the CaEs component, we have here investigated a 106

    series of omphacites from UHP eclogites by FTIR spectroscopy and microprobe analysis, as 107

    well as Mössbauer spectroscopy to fully characterize the cation contents of the structural 108

    formulae. 109

    110

    Geological background and petrography 111

    112

    The investigated eclogites come from the Pohorje Mountains in Slovenia, at the south-eastern 113

    margin of the Alps (Fig. 1). The Pohorje Mountains is mainly formed by metamorphic rocks 114

    of continental basement type, predominantly micaschists, gneisses, marbles, and 115

    metaquarzites (Mioč & Žnidarčič, 1977; Kirst et al., 2010), intruded by granodioritic to 116

    tonalitic plutons and pegmatites in the Miocene (Altherr et al., 1995; Fodor et al., 2008; 117

    Trajanova et al., 2008; Uher et al., 2014). Eclogites, partly amphibolitised, occur in the south-118

    eastern part of Pohorje, together with meta-ultramafic rocks of the Slovenska Bistrica 119

    Ultramafic Complex (SBUC; Hinterlechner-Ravnik et al., 1991; Janák et al., 2006; De Hoog 120

    et al., 2009), metasedimentary gneisses and marbles. 121

  • Skogby et al.

    6

    The first evidence for UHP conditions was documented in kyanite-bearing eclogites 122

    (Janák et al., 2004). Subsequently, UHP conditions were determined for garnet peridotites 123

    from SBUC (Janák et al., 2006) and finally, diamond was found in kyanite and garnet-bearing 124

    gneisses (Janák et al., 2015a) hosting UHP eclogites and garnet peridotites. The UHP 125

    metamorphism in the Pohorje area resulted from subduction of the continental crust during 126

    the Late Cretaceous time (Janák et al. 2004), documented by geochronological data (c.100-90 127

    Ma; Thöni, 2002; Miller et al., 2005; Janák et al., 2009). 128

    The studied eclogites are medium to coarse-grained, with macroscopically visible 129

    reddish garnet, pale-green omphacite, bluish kyanite and pale zoisite, partly amphibolitised 130

    with dark-green amphibole. Microscopic observations show that primary minerals – garnet, 131

    omphacite, kyanite and phengite are variably replaced by retrograde ones forming 132

    symplectites, i.e. diopside + plagioclase after omphacite, plagioclase + biotite after phengite 133

    and sapphirine + corundum + spinel + anorthite after kyanite. Polycrystalline quartz 134

    inclusions in garnet, omphacite and kyanite surrounded by radial cracks indicate breakdown 135

    of former coesite. The most striking feature of omphacite is tiny needles and rods of quartz 136

    (Fig. 2). They display a parallel orientation with the c-axis, indicating exsolution from a pre-137

    existing, more silicic clinopyroxene. Geothermobarometry on the peak metamorphic 138

    assemblage garnet + omphacite + kyanite + phengite + quartz/coesite records pressure and 139

    temperature conditions of up to 3.5-3.7 GPa and 800-850 ºC (Vrabec et al., 2012). More 140

    details on the kyanite eclogites from Pohorje can be found in Janák et al. (2004), Sassi et al. 141

    (2004) or Vrabec et al. (2012), including unusually Cr-rich minerals in these eclogites - 142

    kyanite, Mg-staurolite and corundum (Janák et al., 2015b). 143

    The investigated samples come from 6 localities in Pohorje (Fig. 1): Jurišna Vas (JV4, 144

    PO6), Novak (NO1), Tinjska Gora (PO4), Visole (VI04), Vranjekov Vrh (PO1) and Nova 145

    Gora (NG1), see Vrabec et al. (2012) for a more detailed map and sample location. 146

  • Skogby et al.

    7

    147

    Experimental methods 148

    149

    Infrared spectroscopy 150

    All the seven omphacite samples were measured by Fourier transform infrared (FTIR) 151

    absorption spectroscopy to characterize their OH vibrational bands. The spectrometer system 152

    consisted of a Bruker Equinox 55 spectrometer equipped with a halogen lamp source, a CaF2 153

    beam-splitter, a wire-grid polarizer (KRS-5) and an InSb detector. The crystals were oriented 154

    by morphology and optical microscopy and polished on two sides parallel to (100) and (010). 155

    To avoid cracks and turbid regions, the 100 to 200 m thick crystals sections were mounted 156

    on circular 100-200 m apertures under the microscope. Polarized absorption spectra were 157

    then acquired in the three principal optical directions (, and ) in the wavenumber range 158

    2000-5000 cm-1

    with a resolution of 4 cm-1

    . 159

    To obtain absorption areas for the OH bands the measured spectra were fitted using 160

    the software Peakfit. A baseline subtraction was first applied before fitting with four Voight-161

    shaped bands that produced a good fit to the experimental spectra. Water concentrations were 162

    then calculated from the sum of the integrated absorption areas obtained in the , and 163

    directions (Atot=A+A+A) according to Beer’s law using the calibrations of Koch-Müller et 164

    al. (2007), Libowitzky & Rossman (1997) and Bell et al. (1995). 165

    166

    Mössbauer spectroscopy 167

    The samples were measured by Mössbauer spectroscopy in order to determine the oxidation 168

    state of iron, which is necessary for an accurate normalization of the structural formulae. 169

    Since the samples contained ubiquitous inclusions and turbid areas, the point-source 170

    technique (e.g. McCammon, 1994) was used. Small amounts (ca. 1-2 mg) of clear, inclusion-171

  • Skogby et al.

    8

    free crystal fragments were selected under the microscope. The samples were finely ground, 172

    mixed with a thermoplastic resin and shaped to millimetre-sized cylindrical absorbers under 173

    mild heating. Mössbauer spectra were acquired using a conventional spectrometer system 174

    operated in constant-acceleration mode equipped with a 57

    Co point source with a nominal 175

    activity of 10 mCi and an active area of 0.5 x 0.5 mm. The absorbers were mounted on strip 176

    tape and positioned close (< 1 mm) to the tip of the point source. Spectra were collected in 177

    1024 channels over the velocity range -4.5 to 4.5 mms-1

    , and were calibrated against an -iron 178

    foil before folding and reduction to 256 channels. Spectral fitting was performed with the 179

    least-squares program MDA (Jernberg & Sundqvist, 1983), using a fitting model with two 180

    quadrupole doublets for Fe2+

    and one doublet for Fe3+

    . 181

    182

    Electron microprobe analysis 183

    Electron microprobe analyses (EMPA) of the major elements (Al, Ti, Fe, Mg, Na, K, Si, Ca, 184

    Mn, Cr) were performed at the Department of Earth Sciences, Uppsala University using a FE-185

    EPMA JXA-8530F JEOL superprobe. Between 10 to 30 spot analyses were performed on 186

    each sample using a beam current of 10 nA, an acceleration voltage of 15 kV and a beam 187

    diameter of 1μm. A range of natural and synthetic compounds were used as standards: fayalite 188

    (Fe), periclase (Mg), pyrophanite (Mn, Ti) corundum (Al), wollastonite (Ca and Si), eskolaite 189

    (Cr), albite (Na) and orthoclase (K). 190

    191

    Results 192

    193

    The EMPA results show that the studied crystals are homogenous, whereas the compositions 194

    for the different samples show substantial variation (Table 1). The samples have rather typical 195

    omphacite compositions dominated by diopside (47-59 %) and jadeite (25-33 %) components, 196

  • Skogby et al.

    9

    with minor amounts also of hedenbergite (6-7 %), Ca-Tschermak (4-5 %), enstatite (3-4 %) 197

    and CaEs (3-5 %) components. In back-scatter electron images, some samples show the 198

    occurrence of micrometre-sized quartz-rods (Fig. 3), as previously described by Janák et al. 199

    (2004) and Vrabec et al. (2012) in other crystals partly from the same samples. 200

    Polarized FTIR spectra of oriented omphacite crystals show intense absorption bands 201

    around 3470 and 3520 cm-1

    , strongly polarized in the -direction (Fig. 4). Considerably 202

    weaker bands polarized in the and -directions occur around 3620 cm-1

    . For a few 203

    samples (e.g. NO1), additional weak but sharp bands occur around 3670 cm-1

    , indicative of 204

    amphibole inclusions. These amphibole bands are stronger in spectra from regions close to 205

    cracks and grain boundaries, and in turbid regions. By masking off the IR-beam by small 206

    apertures (100-200 m) such regions could be avoided. Using the molar absorptivity of 207

    65 000 L.mol

    -1.cm

    -2 established for omphacite by Koch-Müller et al., 2007, the recorded 208

    spectra correspond to water concentrations in the range 530 – 870 ppm H2O (Table 1). Water 209

    concentrations calculated based on the wavenumber-dependent general calibration of 210

    Libowitzky & Rossman (1997) are in good agreement, with slightly higher values (up to 6 % 211

    higher). However, the calibration by Bell et al. (1995) results in concentrations which are 212

    approximately twice as high. Similar deviations were observed by Koch-Müller et al. (2004) 213

    for a series of omphacites from Yakutian kimberlite pipes. The Bell et al. (1995) calibration 214

    was based on an kimberlitic augite sample, which has considerably stronger bands at the 215

    higher wavenumbers (3620 cm-1

    ) than typical omphacite spectra, and the observed deviation 216

    is probably related to a wavenumber-dependence of the molar absorptivities. For the 217

    evaluation of the experimental data we have used the mineral-specific calibration of Koch-218

    Müller et al., 2007 which we consider to be that at present most accurate. 219

    Mössbauer spectra were obtained by the point-source method for all samples, and 220

    could be fitted satisfactorily by the three-doublet fitting model mentioned above. 221

  • Skogby et al.

    10

    Representative spectra are displayed in Fig. 5. The obtained hyperfine parameters are well in 222

    line with previous studies of omphacite (e.g. Li et al., 2005), and are listed in Table 2, 223

    together with assignments and absorption ratios. The results show that the Fe3+

    /Fetot ratio vary 224

    substantially, from 0.14 to 0.29 for the sample set. In line with Koch-Müller et al. (2007), the 225

    two fitted Fe2+

    doublets can be considered to be caused by Fe2+

    in the M1 site (outer doublet) 226

    and M2 site (inner doublet). However, the lack of resolution for these two doublets introduces 227

    a large degree of uncertainty in the obtained Fe2+

    distribution. In spite of this, the Fe3+

    /Fetot 228

    ratios are well constrained by the asymmetry of the two main high- and low-velocity 229

    components in the measured spectra, as the Fe3+

    contribution occurs entirely within the low-230

    velocity component and are hence not affected by uncertainties in the Fe2+

    distribution over 231

    the M1 and M2 sites. 232

    Structural formulae were calculated based on 12 positive charges taking all type of 233

    cations into account, including the Fe2+

    /Fe3+

    distribution obtained from Mössbauer 234

    spectroscopy and the hydrogen content obtained from FTIR spectroscopy (Table 1). Without 235

    exception, this lead to cation sums less than 4 (excluding H+), which demonstrate the 236

    occurrence of cation vacancies. CaEs (Ca0.5□0.5AlSi2O6) components were then calculated as 237

    CaEs = 2 x vacancies per formula unit, and varied from 2.5 to 4.6 % for the studied samples. 238

    239

    Discussion 240

    241

    In evaluating intrinsic pyroxene water contents based on the intensity of OH-bands in IR-242

    spectra, it is important to identify absorption bands from possible contaminants. These may be 243

    caused by fluid or melt inclusions, as well as inclusions of hydrous minerals. Among hydrous 244

    mineral inclusions, amphiboles have been shown to occur relatively frequently in 245

    clinopyroxenes, particularly in samples from metamorphic rocks (e.g. Veblen & Buseck, 246

  • Skogby et al.

    11

    1981; Skogby et al., 1990). Fortunately, amphibole OH-bands normally have a characteristic 247

    signature consisting of relatively sharp bands in the 3625-3725 cm-1

    region (e.g. Skogby & 248

    Rossman, 1991; Della Ventura et al., 2003), and can hence easily be distinguished from 249

    intrinsic clinopyroxene OH-bands that occur at lower wavenumbers. More problematic, 250

    however, is the possible occurrence of certain nanometer-sized sheet silicates. As 251

    demonstrated by Koch-Müller et al. (2004) using TEM and synchrotron-IR, absorption bands 252

    in the range 3600-3624 cm-1

    of omphacite spectra can be caused by sub-micrometer 253

    inclusions of clinochlore and amesite. In particular, this seems to be the case when strong 254

    bands occur in this range (e.g. Katayama & Nakashima, 2003), and if not identified, may lead 255

    to significant overestimation of the omphacite water contents. Nevertheless, weak bands that 256

    normally occur in the 3620 cm-1

    range of omphacite spectra (Skogby et al., 1990; Smyth et 257

    al., 1991, Konzett et al., 2008a) may still be caused by intrinsic OH. This is not unexpected 258

    since absorption bands in this range, with the same pleochroism (A A), are typical for 259

    diopside as well as augite (e.g. Skogby, 2006) which both occur as significant components in 260

    normal omphacite compositions. Furthermore, Bromiley & Keppler (2004) observed a major 261

    band at 3613 cm-1

    in spectra of synthetic jadeite and Na-rich clinopyroxene which they 262

    interpreted to be caused by OH coupled to M2-site vacancies. In the present study, the 263

    relatively weak bands in the 3620 cm-1

    range were therefore included in the calculation of 264

    omphacite water contents. The contribution from these bands was however minor, amounting 265

    to 20-30 ppm H2O for the studied samples. 266

    In general, the OH contents are expected to increase with the T, Ptot and PH2O 267

    conditions, related to an increased amount of vacancies formed by stabilization of the CaEs 268

    component. An increase in OH contents with equilibration pressure was also found by 269

    Katayama & Nakashima (2003) for a series of clinopyroxenes from the Kokchetav UHP 270

    metamorphic terrane. Experimental studies (e.g. Gasparik, 1986; Konzett et al., 2008b) have 271

  • Skogby et al.

    12

    shown that the CaEs component in omphacite in addition to bulk composition is also largely 272

    controlled by temperature and pressure conditions. However, it should be noted that the OH 273

    content of omphacite is not merely linked to equilibration pressure conditions. In their study 274

    of omphacites from Yakutian kimberlite pipes, Koch-Müller et al. (2004) noted that samples 275

    from the highest pressures environments (i.e. diamond-bearing eclogite xenoliths) showed the 276

    lowest OH contents, which they interpreted to be caused by either low water activities during 277

    crystallisation or, alternatively, to hydrogen loss during the uplift process. 278

    The CaEs component, and hence the amount of vacancies at the structural M2 site, 279

    display a positive correlation with the OH concentration in our samples (Fig. 6). This 280

    observation is in line with previous studies (Smyth et al., 1991; Koch-Müller et al., 2004; 281

    Katayama & Nikashima, 2003). Due to the determination of Fe2+

    /Fe3+

    ratios for the present 282

    study, the underestimation in calculation of the CaEs component, sometimes resulting in 283

    negative values, is here avoided. Although the regression line fitted to the experimental data 284

    points in Fig. 6 is associated with some uncertainty, the intercept with the y-axis (i.e. for CaEs 285

    = 0) indicates that an omphacite without the M2-vacancies may take up approximately 200 286

    ppm H2O, and that the water content thereafter will increase progressively with the amount of 287

    M2 vacancies. If the observed trend-line is extrapolated to the considerably higher CaEs 288

    components of up to 10 mol-% indicated by the composition of primary omphacite estimated 289

    from the modal content of the precipitated quartz rods mentioned above (Janák et al. 2004), 290

    an original water content of ca 1500 ppm H2O is obtained. Due to the rather limited variation 291

    in OH contents and CaEs components for the studied samples, this extrapolation is associated 292

    with substantial uncertainty. However, the estimated content is in the similar range as that of 293

    ca 2000 ppm H2O calculated by Smyth et al. (1991) for clinopyroxenes from South-African 294

    kimberlite samples, based on a vacancy-rich precursor composition reconstructed from 295

    exsolved garnet and kyanite. 296

  • Skogby et al.

    13

    In their study on hydroxyl contents in omphacite and omphacitic clinopyroxenes from 297

    the Siberian platform, Koch-Müller et al. (2004) observed a pronounced correlation between 298

    the absorbance of the OH-band at 3650-3540 cm-1

    and the amount of Al in the tetrahedral site. 299

    Such a correlation is not obvious for our sample series, however, the much more limited 300

    variation in Al contents for our sample set (Table 1) obscures a possible trend. 301

    The OH concentrations here obtained, ranging from 530-870 ppm H2O, are among the 302

    higher reported for omphacite, and somewhat higher than those reported by Konzett et al. 303

    (2008a) for three omphacite samples from Saualpe and Pohorje eclogites. Previous studies 304

    have indeed reported even higher concentrations, but some of these were based on 305

    calibrations that now have become obsolete (Skogby et al., 1990; Smyth et al., 1991) or on 306

    FTIR spectra with OH bands that appear to be associated with hydrous phases (Katayama & 307

    Nakashima, 2003). The high OH concentrations observed for the Pohorje omphacites are in 308

    line with equilibration under hydrous UHP conditions. 309

    310

    Acknowledgements: Financial support from the Swedish Research Council (HS), the Slovak 311

    Research and Development Agency (project APVV-0080-11to MJ) and the Slovak Scientific 312

    Grant Agency VEGA (grant No. 2/0013/12 to MJ) is gratefully acknowledged. 313

    References 314

    Altherr, R., Lugovic, B., Meyer, H.P., Majer, V. (1995): Early Miocene post-collisional calc-315

    alkaline magmatism along the easternmost segment of the periadriatic fault system 316

    (Slovenia and Croatia). Mineral. Petrol., 54, 225–247. 317

  • Skogby et al.

    14

    Andrut, M., Wildner, M., Ingrin, J., Beran, A. (2007): Mechanisms of OH defect 318

    incorporation in naturally occurring hydrothermally formed diopside and jadeite. Phys. 319

    Chem. Minerals, 34, 543-549. 320

    Bell, D.R., Ihinger P.H., Rossman G.R. (1995): Quantitative analysis of trace OH in garnet 321

    and pyroxenes. Am. Mineral., 80, 465–474. 322

    Bromiley, G.D. & Keppler, H. (2004): An experimental investigation of hydroxyl solubility in 323

    jadeite and Na–rich clinopyroxenes. Contrib. Mineral. Petrol., 147, 189–200. 324

    Bolfan-Casanova, N. (2005): Water in Earth’s mantle. Mineral. Mag., 69(3), 229-257. 325

    De Hoog, J.C.M., Janák, M., Vrabec, M., Froitzheim, N. (2009): Serpentinised peridotites 326

    from an ultra-high pressure terrane in the Pohorje Mts. (Eastern Alps, Slovenia): 327

    geochemical constraints on petrogenesis and tectonic setting. Lithos, 109, 209–222. 328

    Della Ventura, G., Hawthorne, F.C., Robert, J.-L., Iezzi, G. (2003): Synthesis and infrared 329

    spectroscopy of amphiboles along the tremolite–pargasite join. Eur. J. Mineral., 15, 341–330

    347. 331

    Fodor, L.I., Gerdes, A., Dunkl, I., Koroknai, B., Pécskay, Z., Trajanova, M., Horváth, P., 332

    Vrabec, M., Jelen, B., Balogh, K., Frisch, W. (2008): Miocene emplacement and rapid 333

    cooling of the Pohorje pluton at the Alpine-Pannonian-Dinaridic junction, Slovenia. 334

    Swiss J. Geosci., 101, 255–271. 335

    Gasparik, T. (1986): Experimental study of subsolidus phase relations and mixing properties 336

    of clinopyroxene in the silica-saturated system CaO-MgO-Al2O3-SiO2. Amer. Mineral., 337

    71, 686-693. 338

  • Skogby et al.

    15

    Hinterlechner-Ravnik, A., Sassi, F.P., Visona, D. (1991): The Austridic eclogites, metabasites 339

    and metaultrabasites from the Pohorje area (eastern Alps, Yugoslavia): 2. The 340

    metabasites and metaultrabasites, and concluding considerations. Rend. Fis. Acc. Lincei, 341

    2, 175–190. 342

    Ingrin, J. & Skogby, H. (2000): Hydrogen in nominally anhydrous upper-mantle minerals: 343

    concentration levels and implications. Eur. J. Mineral., 12, 543–570. 344

    Janák, M., Cornell, D., Froitzheim, N., De Hoog, J.C.M., Broska, I., Vrabec, M., Hurai, V. 345

    (2009): Eclogite-hosting metapelites from the Pohorje Mountains (Eastern Alps): P-T 346

    evolution, zircon geochronology and tectonic implications. Eur. J. Mineral., 21, 1191–347

    1212. 348

    Janák, M., Froitzheim, N., Lupták, B., Vrabec, M., Krogh Ravna, E.J. (2004): First evidence 349

    for ultrahigh-pressure metamorphism of eclogites in Pohorje, Slovenia: Tracing deep 350

    continental subduction in the Eastern Alps. Tectonics. 23, TC5014. 351

    doi:101029/2004TC001641. 352

    Janák, M., Froitzheim, N., Vrabec, M., Krogh Ravna, E.J., Hoog, J.C.M. (2006): 353

    Ultrahighpressure metamorphism and exhumation of garnet peridotite in Pohorje, Eastern 354

    Alps. J. Metam. Geol., 24, 19–31. 355

    Janák, M., Froitzheim, N., Yoshida, K., Sasinková, V., Nosko, M., Kobayashi, T., Hirajima, 356

    T., Vrabec, M. (2015a): Diamond in metasedimentary crustal rocks from Pohorje, Eastern 357

    Alps: a window to deep continental subduction. J. Metam. Geol., 33, 495-512, 358

    doi:10.1111/jmg.12130. 359

    Janák, M., Uher, P., Ravna, E.J.K., Kullerud, K., Vrabec, M. (2015b): Chromium-rich 360

    kyanite, magnesiostaurolite and corundum in ultrahigh-pressure eclogites (examples from 361

  • Skogby et al.

    16

    Pohorje Mountains, Slovenia and Tromsø Nappe, Norway). Eur. J. Mineral., 27, 377-362

    392. 363

    Jernberg, P. & Sundqvist, T. (1983): A versatile Mössbauer analysis program. Uppsala 364

    University, Institute of Physics (UUIP-1090). 365

    Katayama, I. & Nakashima, S. (2003): Hydroxyl in clinopyroxene from the deep subducted 366

    crust: evidence for H2O transport into the mantle. Amer. Mineral., 88, 229–234. 367

    Katayama, I., Nakashima, S., Yurimoto, H. (2006): Water content in natural eclogite and 368

    implication for water transport into the deep upper mantle. Lithos, 86, 245–259. 369

    Kirst, F., Sandmann, S., Nagel, T.J., Froitzheim, N., Janák, M. (2010): Tectonic evolution of 370

    the southeastern part of the Pohorje Mountains (Eastern Alps, Slovenia). Geol. Carpath., 371

    61, 451–461. 372

    Koch-Müller, M., Matsyuk, S.S., Wirth, R. (2004): Hydroxyl in omphacites and omphacitic 373

    clinopyroxenes of upper mantle to lower crustal origin beneath the Siberian platform. 374

    Amer. Mineral., 89, 921–931. 375

    Koch-Müller, M., Abs-Wurmbach, I., Rhede, D., Kahlenberg, V., Matsyuk, S. (2007): 376

    Dehydration experiments on natural omphacites: qualitative and quantitative 377

    characterization by various spectroscopic methods. Phys. Chem. Minerals, 34, 663–678. 378

    Konzett, J., Frost, D.J., Proyer, A., Ulmer, P. (2008b): The Ca-Eskola component in eclogitic 379

    clinopyroxene as a function of pressure, temperature and bulk composition: an 380

    experimental study to 15 GPa with possible implications for the formation of oriented 381

    SiO2-inclusions in omphacite. Contrib. Mineral. Petrol., 155, 215–228. 382

  • Skogby et al.

    17

    Konzett, J., Libowitzky, E., Hejny, C., Miller, C., Zanetti, A. (2008a): Oriented quartz+calcic 383

    amphibole inclusions in omphacite from the Saualpe and Pohorje Mountain eclogites, 384

    Eastern Alps — An assessment of possible formation mechanisms based on IR- and 385

    mineral chemical data and water storage in Eastern Alpine eclogites. Lithos, 106, 336-386

    350. 387

    Li, Y.L., Zheng, Y.F., Fu, B. (2005): Mössbauer spectroscopy of omphacite and garnet pairs 388

    from eclogites; application to geothermobarometry. Amer. Mineral., 90, 90–100. 389

    Libowitzky, E. & Rossman, G.R. (1997): An IR absorption calibration for water in minerals. 390

    Amer Mineral, 82, 1111–1115. 391

    McCammon, C.A. (1994): A Mössbauer milliprobe: Practical considerations. Hyperf. 392

    Interact. 92, 1235-1239. 393

    Miller, C., Mundil, R., Thöni, M., Konzett, J. (2005): Refining the timing of eclogite 394

    metamorphism: a geochemical, petrological, Sm-Nd and U-Pb case study from the 395

    Pohorje Mountains, Slovenia (Eastern Alps). Contrib. Mineral. Petrol., 150, 70–84. 396

    Mioč, P. & Žnidarčič, M. (1977): Geological map of SFRJ 1:100 000, Sheet Slovenj Gradec. 397

    Geological Survey, Ljubljana, Federal Geological Survey, Beograd. 398

    O’Leary, J.A., Gaetani, G.A., Hauri, E.H. (2010): The effect of tetrahedral Al3+

    on the 399

    partitioning of water between clinopyroxene and silicate melt. Earth. Planet. Sci. Lett., 400

    297, 111-120. 401

    Peslier, A.H. (2010): A review of water contents of nominally anhydrous natural minerals in 402

    the mantles of Earth, Mars and the Moon. J. Volcan. Geotherm. Res., 197, 239-258. 403

  • Skogby et al.

    18

    Sassi, R., Mazzolli, C., Miller, C., Konzett, J. (2004): Geochemistry and metamorphic 404

    evolution of the Pohorje Mountain eclogites from the easternmost Austroalpine basement 405

    of the Eastern Alps (Northern Slovenia). Lithos, 78, 235–261. 406

    Skogby, H. (2006): Water in natural mantle minerals I: pyroxenes. Rev. Mineral. Geochem., 407

    62, 155–168. 408

    Skogby, H. & Rossman, G.R. (1991): The intensity of amphibole OH bands in the infrared 409

    absorption spectrum. Phys. Chem. Mineral., 18, 64–68. 410

    Skogby, H., Bell, D.R., Rossman, G.R. (1990): Hydroxide in pyroxene: variations in the 411

    natural environment. Amer. Mineral., 75, 764–774. 412

    Smyth, J.R., Bell, D.R., Rossman, G.R. (1991) Incorporation of hydroxyl in upper-mantle 413

    clinopyroxenes. Nature, 351, 732–735. 414

    Thöni, M.( 2002): Sm-Nd isotope systematics in garnet from different lithologies (Eastern 415

    Alps): age results and an evaluation of potential problems for garnet Sm-Nd 416

    chronometry. Chem. Geol., 185, 255–281. 417

    Trajanova, M., Pécskay, Z., Itaya, T. (2008): K-Ar geochronology and petrography of the 418

    Miocene Pohorje Mountains batholith (Slovenia). Geol. Carpath., 59, 247–260. 419

    Uher, P., Janák, M., Konečný, P., Vrabec, M. (2014): Rare-element granitic pegmatite of 420

    Miocene age emplaced in UHP rocks from Visole, Pohorje Mountains (Eastern Alps, 421

    Slovenia): accessory minerals, monazite and uraninite chemical dating. Geol. Carpath., 422

    65, 131–146. 423

  • Skogby et al.

    19

    Veblen, D.R. & Buseck, P.R. (1981): Hydrous pyriboles and sheet silicates in pyroxenes and 424

    uralites: Intergrown microstructures and reaction mechanisms. Amer. Mineral., 66, 1107-425

    1134. 426

    Vrabec, M., Janák, M., Froitzheim, N., De Hoog, J.C.M. (2012): Phase relations during peak 427

    metamorphism and decompression of the UHP kyanite eclogites, Pohorje Mountains 428

    (Eastern Alps, Slovenia). Lithos, 144-145, 40-55. 429

    Wade, J.A., Plank, T., Hauri, E.H., Kelley, K.A., Roggensack, K., Zimmer, M. (2008): 430

    Prediction of magmatic water contents via measurement of H2O in clinopyroxene 431

    phenocrysts, Geology, 36, 799-802. 432

    Weis, F.A., Skogby, H., Troll, V.R., Deegan F.M., Dahren, B. (2015): Magmatic water 433

    contents determined through clinopyroxene: Examples from the Western Canary Islands, 434

    Spain. Geochem., Geophys., Geosys., 16, 2127–2146. 435

    436

  • Skogby et al.

    20

    Title of tables 437

    Table 1. Chemical composition of omphacite based on microprobe and FTIR analyses. 438

    Table 2. Mössbauer hyperfine parameters for omphacite. 439

    440

  • Skogby et al.

    21

    Table 1. Chemical composition of omphacite based on microprobe and FTIR analyses. 441

    442

    Fe valence ratios determined by Mössbauer spectroscopy. 443

    444

    445

    NO1 PO6 JV4 PO4 VI04 NG1 PO1

    Wt-% oxides SiO2 55.19 55.36 54.87 55.52 55.27 55.42 55.35 Al2O3 7.78 8.36 8.33 9.76 8.84 8.76 9.24 TiO2 0.13 0.13 0.12 0.13 0.12 0.13 0.12 Cr2O3 0.28 0.05 0.11 0.02 0.18 0.13 0.15 MgO 12.39 11.75 11.63 10.56 11.19 11.47 10.98 FeO 1.69 2.00 1.93 2.01 1.86 2.04 1.86 Fe2O3 0.47 0.60 0.65 0.79 0.83 0.38 0.69 MnO 0.02 0.03 0.02 0.07 0.03 0.02 0.02 CaO 18.91 18.28 18.08 16.53 17.50 18.01 17.45 Na2O 3.59 3.99 3.87 4.77 4.36 4.04 4.41 H2O 0.0535 0.0549 0.0636 0.0702 0.0769 0.0854 0.0866 Total 100.52 100.60 99.67 100.21 100.26 100.49 100.37 Structural formulae normalised to 12 charges Si 1.952 1.954 1.954 1.957 1.954 1.955 1.952 Al 0.324 0.348 0.349 0.406 0.368 0.364 0.384 Ti 0.004 0.004 0.004 0.004 0.004 0.004 0.004 Cr 0.008 0.001 0.003 0.001 0.005 0.004 0.004 Mg 0.654 0.618 0.617 0.555 0.589 0.603 0.578 Fe2+ 0.065 0.076 0.074 0.077 0.071 0.078 0.070 Fe3+ 0.016 0.020 0.022 0.027 0.028 0.013 0.024 Mn 0.001 0.001 0.001 0.003 0.001 0.001 0.001 Ca 0.717 0.691 0.690 0.624 0.663 0.681 0.659 Na 0.247 0.273 0.267 0.325 0.299 0.276 0.302 Total 3.986 3.987 3.981 3.977 3.982 3.978 3.978 H2O (ppm) 535 549 636 702 769 854 866 End-member components Di 0.586 0.549 0.543 0.474 0.515 0.537 0.510 Jd 0.247 0.274 0.268 0.327 0.300 0.278 0.303 Hd 0.058 0.068 0.065 0.065 0.062 0.069 0.062 CaTs 0.048 0.046 0.046 0.043 0.046 0.045 0.048 En 0.034 0.033 0.039 0.045 0.034 0.038 0.034 Ca-Esk 0.027 0.025 0.038 0.046 0.036 0.043 0.043

  • Skogby et al.

    22

    Table 2. Mössbauer hyperfine parameters for omphacite. 446

    Sample Assignment cs dq fwhm Int 447

    NO1 Fe2+ (1) 1.17 2.49 0.51 36.9 448

    Fe2+ (2) 1.14 1.95 0.50 43.0 449

    Fe3+ 0.41 0.43 0.55 20.1 450

    451

    PO6 Fe2+ (1) 1.16 2.56 0.50 30.6 452

    Fe2+ (2) 1.13 2.01 0.51 48.2 453

    Fe3+ 0.41 0.40 0.52 21.2 454

    455

    JV4 Fe2+ (1) 1.17 2.58 0.53 30.4 456

    Fe2+ (2) 1.12 2.04 0.52 46.4 457

    Fe3+ 0.42 0.44 0.51 23.3 458

    459

    PO4 Fe2+ (1) 1.14 2.59 0.52 36.3 460

    Fe2+ (2) 1.17 1.94 0.51 37.6 461

    Fe3+ 0.41 0.46 0.52 26.1 462

    463

    VI04 Fe2+ (1) 1.17 2.58 0.52 32.7 464

    Fe2+ (2) 1.14 2.00 0.53 38.6 465

    Fe3+ 0.40 0.42 0.62 28.7 466

    467

    NG1 Fe2+ (1) 1.16 2.61 0.50 31.4 468

    Fe2+ (2) 1.15 1.97 0.51 54.2 469

    Fe3+ 0.42 0.41 0.53 14.5 470

    471

    PO1 Fe2+ (1) 1.16 2.63 0.58 34.5 472

    Fe2+ (2) 1.14 1.98 0.57 40.4 473

    Fe3+ 0.39 0.47 0.59 25.1 474

    cs=centroid shift, dq=quadrupole splitting, fwhm=full 475

    width at half maximum, Int=percentage of total absorption area. 476

    477

    478

  • Skogby et al.

    23

    Figure captions 479

    480

    Fig. 1. Simplified geological map of Pohorje and adjacent areas (modified from Mioč & 481

    Žnidarčič 1997), showing location of investigated rocks (white stars) near Slovenska Bistrica. 482

    483

    Fig. 2. Microphotograph of omphacite (Omp) with quartz (Qz) rods and inclusion surrounded 484

    by radial cracks. Optical microscope, transmitted light, crossed polars. Sample PO6. 485

    486

    Fig. 3. Back-scatter electron image of omphacite (sample PO6) showing crystallographically 487

    oriented quartz-rods. Length of arrow 20 m. 488

    489

    Fig. 4. Polarized FTIR spectra of omphacite normalized to 1 mm thickness. Spectra are 490

    vertically off-set for clarity. Sinusoidal shortwave variations are interference fringes. Note 491

    weak bands around 3670 cm-1

    in the and directions for sample NO1 caused by amphibole 492

    inclusions. 493

    494

    Fig. 5. Room-temperature Mössbauer spectra of omphacite with fitted sub-spectra. Blue and 495

    red sub-spectra are assigned to Fe2+

    , green sub-spectra are assigned to Fe3+

    . 496

    497

    Fig. 6. Plot of omphacite water content as a function of Ca-Eskola (Ca0.5□0.5AlSi2O6) 498

    component. 499

    500

  • Skogby et al.

    24

    501

    Fig. 1 502

  • Skogby et al.

    25

    503

    Fig. 2 504

    505

    506

    Fig. 3. 507

    508

  • Skogby et al.

    26

    509

    510

    Fig. 4. 511

    512

    513

  • Skogby et al.

    27

    514

    Fig. 5. 515

  • Skogby et al.

    28

    516

    517

    Fig. 6. 518

    519